Magnetochemistry

Last updated

Magnetochemistry is concerned with the magnetic properties of chemical compounds. Magnetic properties arise from the spin and orbital angular momentum of the electrons contained in a compound. Compounds are diamagnetic when they contain no unpaired electrons. Molecular compounds that contain one or more unpaired electrons are paramagnetic. The magnitude of the paramagnetism is expressed as an effective magnetic moment, μeff. For first-row transition metals the magnitude of μeff is, to a first approximation, a simple function of the number of unpaired electrons, the spin-only formula. In general, spin–orbit coupling causes μeff to deviate from the spin-only formula. For the heavier transition metals, lanthanides and actinides, spin–orbit coupling cannot be ignored. Exchange interaction can occur in clusters and infinite lattices, resulting in ferromagnetism, antiferromagnetism or ferrimagnetism depending on the relative orientations of the individual spins.

Contents

Magnetic susceptibility

The primary measurement in magnetochemistry is magnetic susceptibility. This measures the strength of interaction on placing the substance in a magnetic field. The volume magnetic susceptibility, represented by the symbol is defined by the relationship

where, is the magnetization of the material (the magnetic dipole moment per unit volume), measured in amperes per meter (SI units), and is the magnetic field strength, also measured in amperes per meter. Susceptibility is a dimensionless quantity. For chemical applications the molar magnetic susceptibilitymol) is the preferred quantity. It is measured in m3·mol−1 (SI) or cm3·mol−1 (CGS) and is defined as

where ρ is the density in kg·m−3 (SI) or g·cm−3 (CGS) and M is molar mass in kg·mol−1 (SI) or g·mol−1 (CGS).

Schematic diagram of Gouy balance Gouy bal.png
Schematic diagram of Gouy balance

A variety of methods are available for the measurement of magnetic susceptibility.

Types of magnetic behaviour

When an isolated atom is placed in a magnetic field there is an interaction because each electron in the atom behaves like a magnet, that is, the electron has a magnetic moment. There are two types of interaction.

  1. Diamagnetism. When placed in a magnetic field the atom becomes magnetically polarized, that is, it develops an induced magnetic moment. The force of the interaction tends to push the atom out of the magnetic field. By convention diamagnetic susceptibility is given a negative sign. Very frequently diamagnetic atoms have no unpaired electrons ie each electron is paired with another electron in the same atomic orbital. The moments of the two electrons cancel each other out, so the atom has no net magnetic moment. However, for the ion Eu3+ which has six unpaired electrons, the orbital angular momentum cancels out the electron angular momentum, and this ion is diamagnetic at zero Kelvin.
  2. Paramagnetism. At least one electron is not paired with another. The atom has a permanent magnetic moment. When placed into a magnetic field, the atom is attracted into the field. By convention paramagnetic susceptibility is given a positive sign.

When the atom is present in a chemical compound its magnetic behaviour is modified by its chemical environment. Measurement of the magnetic moment can give useful chemical information.

In certain crystalline materials individual magnetic moments may be aligned with each other (magnetic moment has both magnitude and direction). This gives rise to ferromagnetism, antiferromagnetism or ferrimagnetism. These are properties of the crystal as a whole, of little bearing on chemical properties.

Diamagnetism

Diamagnetism is a universal property of chemical compounds, because all chemical compounds contain electron pairs. A compound in which there are no unpaired electrons is said to be diamagnetic. The effect is weak because it depends on the magnitude of the induced magnetic moment. It depends on the number of electron pairs and the chemical nature of the atoms to which they belong. This means that the effects are additive, and a table of "diamagnetic contributions", or Pascal's constants, can be put together. [6] [7] [8] With paramagnetic compounds the observed susceptibility can be adjusted by adding to it the so-called diamagnetic correction, which is the diamagnetic susceptibility calculated with the values from the table. [9]

Paramagnetism

Mechanism and temperature dependence

Variation of magnetic susceptibility with temperature Susceptibility.png
Variation of magnetic susceptibility with temperature

A metal ion with a single unpaired electron, such as Cu2+, in a coordination complex provides the simplest illustration of the mechanism of paramagnetism. The individual metal ions are kept far apart by the ligands, so that there is no magnetic interaction between them. The system is said to be magnetically dilute. The magnetic dipoles of the atoms point in random directions. When a magnetic field is applied, first-order Zeeman splitting occurs. Atoms with spins aligned to the field slightly outnumber the atoms with non-aligned spins. In the first-order Zeeman effect the energy difference between the two states is proportional to the applied field strength. Denoting the energy difference as ΔE, the Boltzmann distribution gives the ratio of the two populations as , where k is the Boltzmann constant and T is the temperature in kelvins. In most cases ΔE is much smaller than kT and the exponential can be expanded as 1 – ΔE/kT. It follows from the presence of 1/T in this expression that the susceptibility is inversely proportional to temperature. [10]

This is known as the Curie law and the proportionality constant, C, is known as the Curie constant, whose value, for molar susceptibility, is calculated as [11]

where N is the Avogadro constant, g is the Landé g-factor, and μB is the Bohr magneton. In this treatment it has been assumed that the electronic ground state is not degenerate, that the magnetic susceptibility is due only to electron spin and that only the ground state is thermally populated.

While some substances obey the Curie law, others obey the Curie-Weiss law.

Tc is the Curie temperature. The Curie-Weiss law will apply only when the temperature is well above the Curie temperature. At temperatures below the Curie temperature the substance may become ferromagnetic. More complicated behaviour is observed with the heavier transition elements.

Effective magnetic moment

When the Curie law is obeyed, the product of molar susceptibility and temperature is a constant. The effective magnetic moment, μeff is then defined [12] as

Where C has CGS units cm3 mol−1 K, μeff is

Where C has SI units m3 mol−1 K, μeff is

The quantity μeff is effectively dimensionless, but is often stated as in units of Bohr magnetonB). [12]

For substances that obey the Curie law, the effective magnetic moment is independent of temperature. For other substances μeff is temperature dependent, but the dependence is small if the Curie-Weiss law holds and the Curie temperature is low.

Temperature independent paramagnetism

Compounds which are expected to be diamagnetic may exhibit this kind of weak paramagnetism. It arises from a second-order Zeeman effect in which additional splitting, proportional to the square of the field strength, occurs. It is difficult to observe as the compound inevitably also interacts with the magnetic field in the diamagnetic sense. Nevertheless, data are available for the permanganate ion. [13] It is easier to observe in compounds of the heavier elements, such as uranyl compounds.

Exchange interactions

Copper(II) acetate dihydrate Copper(II)-acetate-3D-balls2.png
Copper(II) acetate dihydrate
Ferrimagnetic ordering in 2 dimensions Ferrimagnetic ordering.svg
Ferrimagnetic ordering in 2 dimensions
Antiferromagnetic ordering in 2 dimensions Antiferromagnetic ordering.svg
Antiferromagnetic ordering in 2 dimensions

Exchange interactions occur when the substance is not magnetically dilute and there are interactions between individual magnetic centres. One of the simplest systems to exhibit the result of exchange interactions is crystalline copper(II) acetate, Cu2(OAc)4(H2O)2. As the formula indicates, it contains two copper(II) ions. The Cu2+ ions are held together by four acetate ligands, each of which binds to both copper ions. Each Cu2+ ion has a d9 electronic configuration, and so should have one unpaired electron. If there were a covalent bond between the copper ions, the electrons would pair up and the compound would be diamagnetic. Instead, there is an exchange interaction in which the spins of the unpaired electrons become partially aligned to each other. In fact two states are created, one with spins parallel and the other with spins opposed. The energy difference between the two states is so small their populations vary significantly with temperature. In consequence the magnetic moment varies with temperature in a sigmoidal pattern. The state with spins opposed has lower energy, so the interaction can be classed as antiferromagnetic in this case. [14] It is believed that this is an example of superexchange, mediated by the oxygen and carbon atoms of the acetate ligands. [15] Other dimers and clusters exhibit exchange behaviour. [16]

Exchange interactions can act over infinite chains in one dimension, planes in two dimensions or over a whole crystal in three dimensions. These are examples of long-range magnetic ordering. They give rise to ferromagnetism, antiferromagnetism or ferrimagnetism, depending on the nature and relative orientations of the individual spins. [17]

Compounds at temperatures below the Curie temperature exhibit long-range magnetic order in the form of ferromagnetism. Another critical temperature is the Néel temperature, below which antiferromagnetism occurs. The hexahydrate of nickel chloride, NiCl2·6H2O, has a Néel temperature of 8.3 K. The susceptibility is a maximum at this temperature. Below the Néel temperature the susceptibility decreases and the substance becomes antiferromagnetic. [18]

Complexes of transition metal ions

The effective magnetic moment for a compound containing a transition metal ion with one or more unpaired electrons depends on the total orbital and spin angular momentum of the unpaired electrons, and , respectively. "Total" in this context means "vector sum". In the approximation that the electronic states of the metal ions are determined by Russell-Saunders coupling and that spin–orbit coupling is negligible, the magnetic moment is given by [19]

Spin-only formula

Orbital angular momentum is generated when an electron in an orbital of a degenerate set of orbitals is moved to another orbital in the set by rotation. In complexes of low symmetry certain rotations are not possible. In that case the orbital angular momentum is said to be "quenched" and is smaller than might be expected (partial quenching), or zero (complete quenching). There is complete quenching in the following cases. Note that an electron in a degenerate pair of dx2–y2 or dz2 orbitals cannot rotate into the other orbital because of symmetry. [20]

Quenched orbital angular momentum
dnOctahedralTetrahedral
high-spinlow-spin
d1e1
d2e2
d3t2g3
d4t2g3eg1
d5t2g3eg2
d6t2g6e3t23
d7t2g6eg1e4t23
d8t2g6eg2
d9t2g6eg3
legend: t2g, t2 = (dxy, dxz, dyz). eg, e = (dx2–y2, dz2).

When orbital angular momentum is completely quenched, and the paramagnetism can be attributed to electron spin alone. The total spin angular momentum is simply half the number of unpaired electrons and the spin-only formula results.

where n is the number of unpaired electrons. The spin-only formula is a good first approximation for high-spin complexes of first-row transition metals. [21]

IonNumber of
unpaired
electrons
Spin-only
moment /μB
observed
moment /μB
Ti3+11.731.73
V4+11.731.68–1.78
Cu2+11.731.70–2.20
V3+22.832.75–2.85
Ni2+22.832.8–3.5
V2+33.873.80–3.90
Cr3+33.873.70–3.90
Co2+33.874.3–5.0
Mn4+33.873.80–4.0
Cr2+44.904.75–4.90
Fe2+44.905.1–5.7
Mn2+55.925.65–6.10
Fe3+55.925.7–6.0

The small deviations from the spin-only formula may result from the neglect of orbital angular momentum or of spin–orbit coupling. For example, tetrahedral d3, d4, d8 and d9 complexes tend to show larger deviations from the spin-only formula than octahedral complexes of the same ion, because "quenching" of the orbital contribution is less effective in the tetrahedral case. [22]

Low-spin complexes

Crystal field diagram for octahedral low-spin d CFT-Low Spin Splitting Diagram-Vector.svg
Crystal field diagram for octahedral low-spin d
Crystal field diagram for octahedral high-spin d CFT-High Spin Splitting Diagram-Vector.svg
Crystal field diagram for octahedral high-spin d

According to crystal field theory, the d orbitals of a transition metal ion in an octahedal complex are split into two groups in a crystal field. If the splitting is large enough to overcome the energy needed to place electrons in the same orbital, with opposite spin, a low-spin complex will result.

High and low -spin octahedral complexes
d-countNumber of unpaired electronsexamples
high-spinlow-spin
d442Cr2+, Mn3+
d551Mn2+, Fe3+
d640Fe2+, Co3+
d731Co2+

With one unpaired electron μeff values range from 1.8 to 2.5 μB and with two unpaired electrons the range is 3.18 to 3.3 μB. Note that low-spin complexes of Fe2+ and Co3+ are diamagnetic. Another group of complexes that are diamagnetic are square-planar complexes of d8 ions such as Ni2+ and Rh+ and Au3+.

Spin cross-over

When the energy difference between the high-spin and low-spin states is comparable to kT (k is the Boltzmann constant and T the temperature) an equilibrium is established between the spin states, involving what have been called "electronic isomers". Tris-dithiocarbamato iron(III), Fe(S2CNR2)3, is a well-documented example. The effective moment varies from a typical d5 low-spin value of 2.25 μB at 80 K to more than 4 μB above 300 K. [23]

2nd and 3rd row transition metals

Crystal field splitting is larger for complexes of the heavier transition metals than for the transition metals discussed above. A consequence of this is that low-spin complexes are much more common. Spin–orbit coupling constants, ζ, are also larger and cannot be ignored, even in elementary treatments. The magnetic behaviour has been summarized, as below, together with an extensive table of data. [24]

d-countkT/ζ=0.1
μeff
kT/ζ=0
μeff
Behaviour with large spin–orbit coupling constant, ζnd
d10.630μeff varies with T1/2
d21.551.22μeff varies with T, approximately
d33.883.88Independent of temperature
d42.640μeff varies with T1/2
d51.951.73μeff varies with T, approximately

Lanthanides and actinides

Russell-Saunders coupling, LS coupling, applies to the lanthanide ions, crystal field effects can be ignored, but spin–orbit coupling is not negligible. Consequently, spin and orbital angular momenta have to be combined

and the calculated magnetic moment is given by

Magnetic properties of trivalent lanthanide compounds [25]
lanthanideCePrNdPmSmEuGdTbDyHoErTmYbLu
Number of unpaired électrons12345676543210
calculated moment /μB2.543.583.622.680.8507.949.7210.6510.69.587.564.540
observed moment /μB2.3–2.53.4–3.63.5–3.61.4–1.73.3–3.57.9–8.09.5–9.810.4–10.610.4–10.79.4–9.67.1–7.54.3–4.90

In actinides spin–orbit coupling is strong and the coupling approximates to jj coupling.

This means that it is difficult to calculate the effective moment. For example, uranium(IV), f2, in the complex [UCl6]2− has a measured effective moment of 2.2 μB, which includes a contribution from temperature-independent paramagnetism. [26]

Main group elements and organic compounds

Simulated EPR spectrum of the CH3* radical EPR methyl.png
Simulated EPR spectrum of the CH3• radical
MSTL spin-label MTSL chemical structure.png
MSTL spin-label

Very few compounds of main group elements are paramagnetic. Notable examples include: oxygen, O2; nitric oxide, NO; nitrogen dioxide, NO2 and chlorine dioxide, ClO2. In organic chemistry, compounds with an unpaired electron are said to be free radicals. Free radicals, with some exceptions, are short-lived because one free radical will react rapidly with another, so their magnetic properties are difficult to study. However, if the radicals are well separated from each other in a dilute solution in a solid matrix, at low temperature, they can be studied by electron paramagnetic resonance (EPR). Such radicals are generated by irradiation. Extensive EPR studies have revealed much about electron delocalization in free radicals. The simulated spectrum of the CH3• radical shows hyperfine splitting due to the interaction of the electron with the 3 equivalent hydrogen nuclei, each of which has a spin of 1/2. [27] [28]

Spin labels are long-lived free radicals which can be inserted into organic molecules so that they can be studied by EPR. [29] For example, the nitroxide MTSL, a functionalized derivative of TEtra Methyl Piperidine Oxide, TEMPO, is used in site-directed spin labeling.

Applications

The gadolinium ion, Gd3+, has the f7 electronic configuration, with all spins parallel. Compounds of the Gd3+ ion are the most suitable for use as a contrast agent for MRI scans. [30] The magnetic moments of gadolinium compounds are larger than those of any transition metal ion. Gadolinium is preferred to other lanthanide ions, some of which have larger effective moments, due to its having a non-degenerate electronic ground state. [31]

For many years the nature of oxyhemoglobin, Hb-O2, was highly controversial. It was found experimentally to be diamagnetic. Deoxy-hemoglobin is generally accepted to be a complex of iron in the +2 oxidation state, that is a d6 system with a high-spin magnetic moment near to the spin-only value of 4.9 μB. It was proposed that the iron is oxidized and the oxygen reduced to superoxide.

Fe(II)Hb (high-spin) + O2 [Fe(III)Hb]O2

Pairing up of electrons from Fe3+ and O2 was then proposed to occur via an exchange mechanism. It has now been shown that in fact the iron(II) changes from high-spin to low-spin when an oxygen molecule donates a pair of electrons to the iron. Whereas in deoxy-hemoglobin the iron atom lies above the plane of the heme, in the low-spin complex the effective ionic radius is reduced and the iron atom lies in the heme plane. [32]

Fe(II)Hb + O2 [Fe(II)Hb]O2 (low-spin)

This information has an important bearing on research to find artificial oxygen carriers.

Compounds of gallium(II) were unknown until quite recently. As the atomic number of gallium is an odd number (31), Ga2+ should have an unpaired electron. It was assumed that it would act as a free radical and have a very short lifetime. The non-existence of Ga(II) compounds was part of the so-called inert-pair effect. When salts of the anion with empirical formula such as [GaCl3] were synthesized they were found to be diamagnetic. This implied the formation of a Ga-Ga bond and a dimeric formula, [Ga2Cl6]2−. [33]

See also

Related Research Articles

<span class="mw-page-title-main">Diamagnetism</span> Ordinary, weak, repulsive magnetism that all materials possess

Diamagnetic materials are repelled by a magnetic field; an applied magnetic field creates an induced magnetic field in them in the opposite direction, causing a repulsive force. In contrast, paramagnetic and ferromagnetic materials are attracted by a magnetic field. Diamagnetism is a quantum mechanical effect that occurs in all materials; when it is the only contribution to the magnetism, the material is called diamagnetic. In paramagnetic and ferromagnetic substances, the weak diamagnetic force is overcome by the attractive force of magnetic dipoles in the material. The magnetic permeability of diamagnetic materials is less than the permeability of vacuum, μ0. In most materials, diamagnetism is a weak effect which can be detected only by sensitive laboratory instruments, but a superconductor acts as a strong diamagnet because it repels a magnetic field entirely from its interior.

<span class="mw-page-title-main">Magnetism</span> Class of physical phenomena

Magnetism is the class of physical attributes that are mediated by a magnetic field, which refers to the capacity to induce attractive and repulsive phenomena in other entities. Electric currents and the magnetic moments of elementary particles give rise to a magnetic field, which acts on other currents and magnetic moments. Magnetism is one aspect of the combined phenomena of electromagnetism. The most familiar effects occur in ferromagnetic materials, which are strongly attracted by magnetic fields and can be magnetized to become permanent magnets, producing magnetic fields themselves. Demagnetizing a magnet is also possible. Only a few substances are ferromagnetic; the most common ones are iron, cobalt, and nickel and their alloys. The rare-earth metals neodymium and samarium are less common examples. The prefix ferro- refers to iron because permanent magnetism was first observed in lodestone, a form of natural iron ore called magnetite, Fe3O4.

<span class="mw-page-title-main">Paramagnetism</span> Weak, attractive magnetism possessed by most elements and some compounds

Paramagnetism is a form of magnetism whereby some materials are weakly attracted by an externally applied magnetic field, and form internal, induced magnetic fields in the direction of the applied magnetic field. In contrast with this behavior, diamagnetic materials are repelled by magnetic fields and form induced magnetic fields in the direction opposite to that of the applied magnetic field. Paramagnetic materials include most chemical elements and some compounds; they have a relative magnetic permeability slightly greater than 1 and hence are attracted to magnetic fields. The magnetic moment induced by the applied field is linear in the field strength and rather weak. It typically requires a sensitive analytical balance to detect the effect and modern measurements on paramagnetic materials are often conducted with a SQUID magnetometer.

<span class="mw-page-title-main">Energy level</span> Different states of quantum systems

A quantum mechanical system or particle that is bound—that is, confined spatially—can only take on certain discrete values of energy, called energy levels. This contrasts with classical particles, which can have any amount of energy. The term is commonly used for the energy levels of the electrons in atoms, ions, or molecules, which are bound by the electric field of the nucleus, but can also refer to energy levels of nuclei or vibrational or rotational energy levels in molecules. The energy spectrum of a system with such discrete energy levels is said to be quantized.

<span class="mw-page-title-main">Curie temperature</span> Temperature above which magnetic properties change

In physics and materials science, the Curie temperature (TC), or Curie point, is the temperature above which certain materials lose their permanent magnetic properties, which can (in most cases) be replaced by induced magnetism. The Curie temperature is named after Pierre Curie, who showed that magnetism was lost at a critical temperature.

<span class="mw-page-title-main">Zeeman effect</span> Spectral line splitting in magnetic field

The Zeeman effect is the effect of splitting of a spectral line into several components in the presence of a static magnetic field. It is named after the Dutch physicist Pieter Zeeman, who discovered it in 1896 and received a Nobel prize for this discovery. It is analogous to the Stark effect, the splitting of a spectral line into several components in the presence of an electric field. Also similar to the Stark effect, transitions between different components have, in general, different intensities, with some being entirely forbidden, as governed by the selection rules.

In electromagnetism, the magnetic susceptibility is a measure of how much a material will become magnetized in an applied magnetic field. It is the ratio of magnetization M to the applied magnetizing field intensity H. This allows a simple classification, into two categories, of most materials' responses to an applied magnetic field: an alignment with the magnetic field, χ > 0, called paramagnetism, or an alignment against the field, χ < 0, called diamagnetism.

<span class="mw-page-title-main">Hyperfine structure</span> Small shifts and splittings in the energy levels of atoms, molecules and ions

In atomic physics, hyperfine structure is defined by small shifts in otherwise degenerate energy levels and the resulting splittings in those energy levels of atoms, molecules, and ions, due to electromagnetic multipole interaction between the nucleus and electron clouds.

<span class="mw-page-title-main">Magnetic moment</span> Magnetic strength and orientation of an object that produces a magnetic field

In electromagnetism, the magnetic moment is the magnetic strength and orientation of a magnet or other object that produces a magnetic field. Examples of objects that have magnetic moments include loops of electric current, permanent magnets, elementary particles, various molecules, and many astronomical objects.

In atomic physics, the spin quantum number is a quantum number which describes the intrinsic angular momentum of an electron or other particle. The phrase was originally used to describe the fourth of a set of quantum numbers, which completely describe the quantum state of an electron in an atom. The name comes from a physical spinning of the electron about an axis, as proposed by Uhlenbeck and Goudsmit. The value of ms is the component of spin angular momentum parallel to a given direction, which can be either +1/2 or –1/2.

The Curie–Weiss law describes the magnetic susceptibility χ of a ferromagnet in the paramagnetic region above the Curie point:

In atomic physics, the electron magnetic moment, or more specifically the electron magnetic dipole moment, is the magnetic moment of an electron resulting from its intrinsic properties of spin and electric charge. The value of the electron magnetic moment is −9.2847647043(28)×10−24 J⋅T−1. The electron magnetic moment has been measured to an accuracy of 1.7×10−13 relative to the Bohr magneton.

In quantum physics, the spin–orbit interaction is a relativistic interaction of a particle's spin with its motion inside a potential. A key example of this phenomenon is the spin–orbit interaction leading to shifts in an electron's atomic energy levels, due to electromagnetic interaction between the electron's magnetic dipole, its orbital motion, and the electrostatic field of the positively charged nucleus. This phenomenon is detectable as a splitting of spectral lines, which can be thought of as a Zeeman effect product of two relativistic effects: the apparent magnetic field seen from the electron perspective and the magnetic moment of the electron associated with its intrinsic spin. A similar effect, due to the relationship between angular momentum and the strong nuclear force, occurs for protons and neutrons moving inside the nucleus, leading to a shift in their energy levels in the nucleus shell model. In the field of spintronics, spin–orbit effects for electrons in semiconductors and other materials are explored for technological applications. The spin–orbit interaction is at the origin of magnetocrystalline anisotropy and the spin Hall effect.

<span class="mw-page-title-main">Electron paramagnetic resonance</span> Technique to study materials that have unpaired electrons

Electron paramagnetic resonance (EPR) or electron spin resonance (ESR) spectroscopy is a method for studying materials that have unpaired electrons. The basic concepts of EPR are analogous to those of nuclear magnetic resonance (NMR), but the spins excited are those of the electrons instead of the atomic nuclei. EPR spectroscopy is particularly useful for studying metal complexes and organic radicals. EPR was first observed in Kazan State University by Soviet physicist Yevgeny Zavoisky in 1944, and was developed independently at the same time by Brebis Bleaney at the University of Oxford.

<span class="mw-page-title-main">Magnetization</span> Physical quantity, density of magnetic moment per volume

In classical electromagnetism, magnetization is the vector field that expresses the density of permanent or induced magnetic dipole moments in a magnetic material. Movement within this field is described by direction and is either Axial or Diametric. The origin of the magnetic moments responsible for magnetization can be either microscopic electric currents resulting from the motion of electrons in atoms, or the spin of the electrons or the nuclei. Net magnetization results from the response of a material to an external magnetic field. Paramagnetic materials have a weak induced magnetization in a magnetic field, which disappears when the magnetic field is removed. Ferromagnetic and ferrimagnetic materials have strong magnetization in a magnetic field, and can be magnetized to have magnetization in the absence of an external field, becoming a permanent magnet. Magnetization is not necessarily uniform within a material, but may vary between different points. Magnetization also describes how a material responds to an applied magnetic field as well as the way the material changes the magnetic field, and can be used to calculate the forces that result from those interactions. It can be compared to electric polarization, which is the measure of the corresponding response of a material to an electric field in electrostatics. Physicists and engineers usually define magnetization as the quantity of magnetic moment per unit volume. It is represented by a pseudovector M.

In chemistry and physics, the exchange interaction is a quantum mechanical effect that only occurs between identical particles. Despite sometimes being called an exchange force in an analogy to classical force, it is not a true force as it lacks a force carrier.

Molecule-based magnets (MBMs) or molecular magnets are a class of materials capable of displaying ferromagnetism and other more complex magnetic phenomena. This class expands the materials properties typically associated with magnets to include low density, transparency, electrical insulation, and low-temperature fabrication, as well as combine magnetic ordering with other properties such as photoresponsiveness. Essentially all of the common magnetic phenomena associated with conventional transition-metal magnets and rare-earth magnets can be found in molecule-based magnets. Prior to 2011, MBMs were seen to exhibit "magnetic ordering with Curie temperature (Tc) exceeding room temperature".

The concept of a double group was introduced by Hans Bethe for the quantitative treatment of magnetochemistry of complexes of ions like Ti3+, that have a single unpaired electron in the metal ion's valence electron shell and to complexes of ions like Cu2+ which have a single "vacancy" in the valence shell.

For many paramagnetic materials, the magnetization of the material is directly proportional to an applied magnetic field, for sufficiently high temperatures and small fields. However, if the material is heated, this proportionality is reduced. For a fixed value of the field, the magnetic susceptibility is inversely proportional to temperature, that is

In condensed matter and atomic physics, Van Vleck paramagnetism refers to a positive and temperature-independent contribution to the magnetic susceptibility of a material, derived from second order corrections to the Zeeman interaction. The quantum mechanical theory was developed by John Hasbrouck Van Vleck between the 1920s and the 1930s to explain the magnetic response of gaseous nitric oxide and of rare-earth salts. Alongside other magnetic effects like Paul Langevin's formulas for paramagnetism and diamagnetism, Van Vleck discovered an additional paramagnetic contribution of the same order as Langevin's diamagnetism. Van Vleck contribution is usually important for systems with one electron short of being half filled and this contribution vanishes for elements with closed shells.

References

  1. Earnshaw, p. 89
  2. Magnetic Susceptibility Balances
  3. O'Connor, C.J. (1982). Lippard, S.J. (ed.). Magnetic susceptibility measurements. Progress in Inorganic Chemistry. Vol. 29. Wiley. p. 203. ISBN   978-0-470-16680-2.
  4. Evans, D.F. (1959). "The determination of the paramagnetic susceptibility of substances in solution by nuclear magnetic resonance". J. Chem. Soc.: 2003–2005. doi:10.1039/JR9590002003.
  5. Orchard, p. 15. Earnshshaw, p. 97
  6. Figgis&Lewis, p. 403
  7. Carlin, p. 3
  8. Bain, Gordon A.; Berry , John F. (2008). "Diamagnetic Corrections and Pascal's Constants". J. Chem. Educ. 85 (4): 532. Bibcode:2008JChEd..85..532B. doi:10.1021/ed085p532.
  9. Figgis&Lewis, p. 417
  10. Figgis&Lewis, p. 419
  11. Orchard, p. 48
  12. 1 2 Hoppe, J.I. (1972). "Effective magnetic moment". J. Chem. Educ. 49 (7): 505. Bibcode:1972JChEd..49..505H. doi:10.1021/ed049p505.
  13. Orchard, p. 53
  14. 1 2 Lawrence Que (March 2000). Physical methods in bioinorganic chemistry: spectroscopy and magnetism. University Science Books. pp. 345–348. ISBN   978-1-891389-02-3 . Retrieved 22 February 2011.
  15. Figgis&Lewis, p. 435. Orchard, p. 67
  16. Carlin, sections 5.5–5.7
  17. Carlin, chapters 6 and 7, pp. 112–225
  18. Carin, p. 264
  19. Figgis&Lewis, p. 420
  20. Figgis&Lewis, pp. 424, 432
  21. Figgis&Lewis, p. 406
  22. Figgis&Lewis, Section 3, "Orbital contribution"
  23. Orchard, p. 125. Carlin, p. 270
  24. Figgis&Lewis, pp. 443–451
  25. Greenwood&Earnshaw p. 1243
  26. Orchard, p. 106
  27. Weil, John A.; Bolton, James R.; Wertz, John E. (1994). Electron paramagnetic resonance : elementary theory and practical applications. Wiley. ISBN   0-471-57234-9.
  28. Atkins, P. W.; Symons, M. C. R. (1967). The structure of inorganic radicals; an application of electron spin resonance to the study of molecular structure. Elsevier.
  29. Berliner, L.J. (1976). Spin labeling : theory and applications I. Academic Press. ISBN   0-12-092350-5.Berliner, L.J. (1979). Spin labeling II : theory and applications. Academic Press. ISBN   0-12-092352-1.
  30. Krause, W. (2002). Contrast Agents I: Magnetic Resonance Imaging: Pt. 1. Springer. ISBN   3540422471.
  31. Caravan, Peter; Ellison, Jeffrey J.; McMurry, Thomas J. ; Lauffer, Randall B., Jeffrey J.; McMurry, Thomas J.; Lauffer, Randall B. (1999). "Gadolinium(III) Chelates as MRI Contrast Agents: Structure, Dynamics, and Applications". Chem. Rev. 99 (9): 2293–2352. doi:10.1021/cr980440x. PMID   11749483.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  32. Greenwood&Earnshaw, pp. 1099–1011
  33. Greenwood&Earnshaw, p. 240

Bibliography